首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 62 毫秒
1.
Electronic structural signatures of the guanine-7H and guanine-9H tautomers have been investigated on an orbital by orbital basis using dual space analysis. A combination of density functional theory (B3LYP/TZVP), the statistical average of model orbital potentials (SAOP/TZ2P) method and outer valence Green's function theory (OVGF/TZVP) has been used to generate optimal tautomer geometries and accurate ionization energy spectra for the guanine tautomer pair. The present work found that the non-planar form for both of the guanine keto pair possesses lower energies than their corresponding planar counterparts, and that the canonical form of the guanine-7H tautomer has slightly lower total energy than guanine-9H. This latter result is in agreement with previous experimental and theoretical findings. In the planar guanine pair the geometric parameters and anisotropic molecular properties are compared, focusing on changes caused by the mobile proton transfer. It is demonstrated that the mobile proton only causes limited disturbance to isotropic properties, such as geometry and the energetics, of the guanine keto tautomer pair. The exception to this general statement is for related local changes such as the N((7))-C((8)) and C((8))-N((9)) bond length resonance between the single and double bonds, reflecting the nitrogen atom being bonded with the mobile proton in the tautomers. The mobile proton distorts the electron distribution of the tautomers, which leads to significant changes in the molecular anisotropic properties. The dipole moment of guanine-7H is altered by about a factor of three, from 2.23 to 7.05 D (guanine-9H), and the molecular electrostatic potentials also reflect significant electron charge distortion. The outer valence orbital momentum distributions, which were obtained using the plane wave impulse approximation (PWIA), have demonstrated quantitatively that the outer valence orbitals of the tautomer pair can be divided into three groups. That is orbitals 1a'-7a' and 18a', which do not have visible alternations in the tautomeric process (which consist of either pi orbitals or are close to the inner valence shell); a second group comprising orbitals 19a'-22a', 25a', 26a', 28a', 29a' and 31a', which show small perturbations as a result of the mobile hydrogen locations; and group three, orbitals 23a', 24a', 27a', 30a' and 32a', which demonstrate significant changes due to the mobile proton transfer and are therefore considered as signature orbitals of the G-7H/G-9H keto tautomeric process.  相似文献   

2.
To test the hypothesis that substrate-induced steric compression between His 57 and Asp 102 at the active site of chymotrypsin can increase the basicity of His 57, we have synthesized thecis- andtrans-isomers of 2-bromo-3-(N-tritylimidazole)-2-propenoic acid and 2-chloro-3-(N-tritylimidazole)-2-propenoic acid and compared selected properties with those ofcis-andtrans-urocanic acids. Thecis-isomers display low field1H NMR signals at 17 ppm in dimethylsulfoxide, similar tocis-urocanic acid; whereas thetrans-isomers do not show strong hydrogen bonds. Increasing the size of the C2 substituent (H < Cl < Br) in thecis-isomers increases the pKaof the imidazolium group from 6.78 for H to 7.81 and 9.10 for Cl and Br, respectively; whereas the pKas of thetransisomers are all 6.0 ± 0.1. The results indicate that thecis-urocanic acid derivatives with large substituents at C2 act as proton sponges in water, and they support the concept that steric compression in the catalytic triad of chymotrypsin can increase the basicity of His 57.  相似文献   

3.
Semi-empirical energy calculations for an internal Pro-Pro dimer are presented that take into account the nature of the flexibility of the proline ring due to its puckering. Calculations show that three stable conformations are available for the dimer: the cis (ω = 0°, ψ = 160°); the trans (ω = 180°, ψ = 160°, also referred to as trans′); and the cis′ (ω = 180°, ψ = ?40°) conformations. The best conformational pathways between these stable conformations are determined. Calculations also show that the barrier for cis′–trans′ conversion is of the same order of magnitude as that for cistrans conversion.  相似文献   

4.
Complexes of the types cis- and trans-Pt(amine)2I2 were studied by spectroscopic methods, especially by multinuclear NMR spectroscopy. In 195Pt NMR, the cis diiodo compounds with primary amines were observed between −3342 and −3357 ppm in acetone, while the trans compounds were found between −3336 and −3372 ppm. For the secondary amines, the chemical shifts were observed at lower fields. In 1H NMR, the trans complexes were observed at higher fields than the cis compounds, while in 13C NMR, the reverse was observed. The 2J(195Pt-1H) and 3J(195Pt-1H) coupling constants are larger for the cis compounds (ave. 67 and 45 Hz, respectively) than for the trans isomers (ave. 59 and 38 Hz). In 13C NMR, the values of 2J(195Pt-13C) and 3J(195Pt-13C) were also found to be larger for the cis complexes (ave. 17 and 39 Hz versus 11 and 28 Hz). There seems to be a slight dependence of the pKa values of the protonated amines or the proton affinity in the gas phase with the δ(Pt) chemical shifts. The crystal structures of eight diiodo complexes were determined. These compounds are cis-Pt(CH3NH2)2I2, cis-Pt(n-C4H9NH2)2I2, cis-Pt(Et2NH)2I2, trans-Pt(n-C3H7NH2)2I2, trans-Pt(iso-C3H7NH2)2I2, trans-Pt(n-C4H9NH2)2I2, trans-Pt(t-C4H9NH2)2I2 and trans-Pt(Me2NH)2I2. The Pt-N bond distances located in trans position to the iodo ligands were compared to those located in trans position to the amines. The Pt-N bond in cis-Pt(Et2NH)2I2 are much longer than the others, probably caused by the steric hindrance of the two very bulky ligands located in cis positions.  相似文献   

5.
Concentration and temperature dependences of the 1H nmr spectra of N-acetyl-L -proline N-methylamide were observed in various solvents [CCl4, CDCl3, (CD3)2CO, (CD3)2SO, H2O, and D2O]. The fraction of the cis isomer (with respect to the bond between the acetyl carbonyl carbon and prolyl nitrogen atoms) depends greatly on the solvent used; the fraction of the cis isomer is higher in polar solvents than in nonpolar solvents. It depends also on concentration and temperature in nonpolar solvents but not in polar solvents. In nonpolar solvents the trans isomer mostly exists in the γ-turn structure with an intramolecular hydrogen bond and the cis isomer tends to form molecular aggregates by intermolecular hydrogen bonds. In polar solvents both the cis and trans isomers exist in monomeric forms which interact with solvent molecules. The pH dependences of the N-methyl proton resonances indicate that the γ-turn structure of the trans isomer is present also in aqueous solution, though its population is difficult to determine. Apparent enthalpy and entropy changes for the conversion of the trans isomer to cis isomer are evaluated for various solvents. The results are discussed in terms of the intra- and intermolecular hydrogen bondings.  相似文献   

6.
Conformations of the cyclic tetrapeptide cyclo(L -Pro-Sar)2 in solution were studied by 1H- and 13C-nmr spectrometry and model building. The nmr data provide definite evidence that this cyclic peptide exists chiefly in two conformations, namely, a C2-symmetric conformation and an asymmetric structure. The former was demonstrated to be predominant in polar solvents (100% in Me2SO-d6). This structure contains all cis-peptide bond linkages and all trans′ Pro Cα?CO bonds. It represents the first cyclic tetrapeptide in which all four peptide bonds have been found in the cis-conformation. As the polarity of the solvent decreases, the population of C2-symmetric conformers decreases (88% in CD3CN and 65% in CDCl3). At the same time, a minor asymmetric conformer, characterized by cis-cis-cis-trans peptide bond sequences (two cis Sar-Pro bonds, one cis Pro-Sar bond, and one trans Pro-Sar bond), is seen to increase (9% in CD3CN and 30% in CDCl3). A proposed predominant conformation in solution for cyclo(L -Pro-Sar)2 was compared with a crystal structure, as reported in an accompanying paper. Both structures show striking overall similarities.  相似文献   

7.
The effects of an electric field and of various substituents on the symmetry breaking of degenerate near-midgap orbitals and on different properties in bi-N,N-pyrazine-1,6-hexatriene dications ([C4N2H4—(CH)6—C4N2H4]2+) are investigated by means of semiempirical PM3 and INDO CI methods. The electric field is simulated by applying positive/negative point charges at varying distances from the end-points, and the substitutions are done with single chlorine atoms or with CN, OH or CH3 groups, at various positions along the chain or on one of the pyrazine rings. The results are compared with calculations on the unsubstituted, field-free system. It is found that an electric field (e.g., as applied over a membrane) leads to significant symmetry breaking and also polarizes the HOMO and LUMO, such that electron transfer between these orbitals generates large dipole-moment shifts and non-negligible oscillator strengths. With substituents, no major symmetry breaking is observed for the ground state. Instead, strong modifications of the orbital picture are observed, in particular when using the stronger electron-withdrawing substituents. Placing the substituent in a ring position does, furthermore, lead to the possibility of large charge transfer.  相似文献   

8.
 The interaction of Fe(II) and Fe(III) with the novel Fe(II) chelator N,N′N″-tris(2-pyridylmethyl)-cis,cis-1,3,5-triaminocyclohexane (referred to as tachpyr) gives rise to six-coordinate, low-spin, cationic complexes of Fe(II). Tachpyr also displays a cytotoxicity toward cultured bladder cancer cells that is believed to involve coordination of intracellular iron. The anaerobic reaction of tachpyr with Fe(II) salts affords the Fe(II)-tachpyr2+ complex, but in presence of oxygen, oxidative dehydrogenation of one or two of the aminomethylene group(s) of the ligand occurs, with formal loss of H2: R—N(H)—C(H)2—(2-py) → R—N=C(H)—(2-py)+H2. The resulting mono- and diimino Fe(II) complexes (denoted as [Fe(tachpyr-H2)]2+ and [Fe(tachpyr-2H2)]2+) are an inseparable mixture, but they may be fully oxidized by H2O2 to the known tris(imino) complex Fe(II)[cis,cis-1,3,5-tris(pyridine-2-carboxaldimino)cyclohexane]2+ (or [Fe(tachpyr-3H2)]2+). Cyclic voltammetry of the imino complex mixture reveals an irreversible anodic wave at +0.78 V vs. NHE. Tachpyr acts as a reducing agent toward Fe(IIII) salts, affording the same two Fe(II) imino complexes as products. Tachpyr also reductively removes Fe(III) from an Fe(III)(ATP)3 complex (which is a putative form of intracellular iron), producing the two Fe(II) imino complexes. Novel N-alkylated derivatives of tachpyr have been synthesized. N-Alkylation has two effects on tachpyr: lowering metal affinity through increased steric hindrance, and preventing Fe(III) reduction because oxidative dehydrogenation of nitrogen is blocked. The N-methyl tachpyr derivative binds Fe(II) only weakly as a high-spin complex, and no complexation or reduction of Fe(III) is observed. Corresponding to their inability to bind iron, the N-alkylated chelators are nontoxic to cultured bladder cancer cells. A tach-based chelator with three N-propyleneamino arms is also synthesized. Studies of the chemical and biochemical properties of this chelator further support a relationship between intracellular iron chelation, iron reduction, and cytotoxicity. Received: 23 March 1998 / Accepted: 1 June 1998  相似文献   

9.
Alkylation of 2-methylthiopyrimidin-4(1H)-one (1a) and its 5(6)-alkyl derivatives 1bd as well as theophylline (7) with 2,2-bis(bromomethyl)-1,3-diacetoxypropane (2) under microwave irradia-tion gave the corresponding acyclonucleosides 1-[(3-acetoxy-2-acetoxymethyl-2-bromomethyl)prop-1-yl]-2-methyl-thio pyrmidin-4(1H)-ones 3ad and 7-[(3-acetoxy-2-acetoxymethyl-2-bromomethyl)prop-1-yl]theophylline (8), which upon further irradiation gave the double-headed acyclonucleosides 1,1 ′-[(2,2-diacetoxymethyl)-1,3-propylidene]-bis[(2-(methylthio)-pyrimidin-4(1H)-ones] 4ac, and 7,7 ′-[(2,2-diacetoxymethyl)-1,3-propylidene]-bis(theophylline) (9). The deacetylated derivatives were obtained by the action of sodium methoxide. The activity of deacetylated nucleosides against Hepatitis B virus was evaluated. Compound 5b showed moderate inhibition activity against HBV with mild cytotoxicity.  相似文献   

10.
The 1,3-oxazine complexes cis- and trans-[PtCl2{ C(R)OCH2CH2C}H22] (cis: R=CH3 (1a), CH2CH3 (2a), (CH3)3C (3a), C6H5 (4a); trans:R =CH3 (1b), C6H5 (4b)) were obtained in 51-71% yield by reaction in THF at 0 °C of the corresponding nitrile complexes cis- and trans-[PtCl2(NCR)2] with 2 equiv. of OCH2CH2CH2Cl, generated by deprotonation of 3-chloro-1-propanol with n-BuLi. The cationic nitrile complexes trans-[Pt(CF3)(NCR)(PPh3)2]BF4 (R=CH3, C6H5) react with 1 equiv, of OCH2CH2CH2Cl to give a mixture of products, including the corresponding oxazine derivatives trans-[Pt(CF3){ CH2}(PPh3)2]BF4 (5 and 6), the chloro complex trans- [Pt(CF3)Cl(PPh3)2] and free oxazine H2. For short reaction times (c. 5–15 min) the oxazine complexes 5 and 6 could be isolated in modest yield (37–49%) from the reaction mixtures and they could be separated from the corresponding chloro complex (yield 40%) by taking advantage of the higher solubility of the latter derivative in benzene. For longer reaction times (> 2 h), trans-[Pt(CF3)Cl(PPh3)2] was the only isolated product. Complex 6 was crystallographically characterized and it was found to contain also crystals of trans- [PtCl{ H2}(PPh3)2]BF4, which prevented a more detailed analysis of the bond lengths and angles within the metal coordination sphere. The 1,3-oxazine ring, which shows an overall planar arrangement, is characterized by high thermal values of the carbon atoms of the methylene groups indicative of disordering in this part of the molecule in agreement with fast dynamic ring processes suggested on the basis of 1H NMR spectra. It crystallizes in the trigonal space group P , with a=22.590(4), b=15.970(3) Å, γ=120°, V=7058(1) Å3 and Z=6. The structure was refined to R=0.059 for 3903 unique observed (I3σ(I)) reflections. A mechanism is proposed for the conversion of nitrile ligands to oxazines in Pt(II) complexes.  相似文献   

11.
 Both cis- and trans-RuCl2(DMSO)4 (cis-Ru and trans-Ru) react with ApG, GpA, d(ApG) and d(GpA) to yield products with bifunctional metal coordination of the bases. For each dinucleotide one major product and several minor species are formed. This is in contrast to previous results on analogous reactions between trans-Ru and d(GpG) where a substantial amount of an intermediate species was found. The rates of reaction between dinucleotides and cis-Ru are approximately 20-fold slower than for trans-Ru. The compounds formed with the two isomers exhibit identical proton NMR spectra, suggesting the same coordination mode for ruthenium in the final product. The two purine bases are coordinated to ruthenium through N7 in a head-to-head conformation with the glycosidic angles being in the anti range. Coupling constants indicate a relatively pure 3′-endo conformation for the 5′-sugar and mainly 2′-endo for the 3′-sugar. The similar bifunctional binding mode of cis- and trans-Ru(II) with dinucleotides as evident from the NMR spectra are in contrast to the different mode of interaction suggested earlier for cis- and trans-Ru complexes with DNA. trans-Ru interacts with the deoxyoctanucleotide d(CCTGGTCC), giving two main products during the first 2 h of incubation time. Four H8 guanine resonances are shifted downfield, characteristic of N7 metal coordination. The products are not analyzed in detail, but it is suggested that the structures may be described as two chiral G(N7/N7) chelates. Received: 20 August 1998 / Accepted: 20 January 1999  相似文献   

12.
1H and 13C high-resolution nmr spectra of cationic, zwitterionic, and anionic forms of the peptides: H-Trp-(Pro)n-Tyr-OH, n = 0-5, and H-Trp-Pro-OCH3 were obtained in D2O solution. Analysis of Hα(Pro1), Hα(Trp), Cγ(Pro), Hε(Tyr), and Hδ(Trp) resonances provided evidence for the presence of two predominant backbone isomers: the all-trans one and another with the Trp-Pro peptide bond in cis conformation; the latter constituted about 0.8 molar fraction of the total peptide (n > 1) concentration. Relative content of these isomers varied in a characteristic way with the number of Pro residues and the ionization state of the peptides. The highest content of the cis (Trp-Pro) isomer, 0.74, was found in the anionic form of H-Trp-Pro-Tyr-OH; it decreased in the order of: anion ? zwitterion ≈ cation, and with the number of Pro residues to reach the value of 0.42 in the cationic form of H-Trp- (Pro)5-Tyr-OH. Isomerization equilibria about Pro-Pro bond(s) were found to be shifted far (?0.9) in favor of the trans conformation. Interpretation of the measured vicinal coupling constants Jα?β′ and Jα?β″ for CαH-CβH2 proton systems of Trp and Tyr side chains in terms of relative populations of g+, g?, and t staggered rotamers around the χ1 dihedral angle indicated that in all the peptides studied (a) rotation of Trp indole ring in cis (Trp-Pro) isomers is strongly restricted, and (b) rotation of Tyr phenol ring is relatively free. The most preferred χ1 rotamer of Trp (0.8-0.9 molar fraction) was assigned as the t one on the basis of a large value of the vicinal coupling constant between the high-field Hβ and carbonyl carbon atoms of Trp, estimated for the cis (Pro1) form of H-Trp-Pro-Tyr-OH from a 1H, 13C correlated spectroscopy 1H detected multiple quantum experiment. This indicates that cis ? trans equilibrium in the Trp-Pro fragment is governed by nonbonding interactions between the pyrrolidine (Pro) and indole (Trp) rings. A molecular model of the terminal cis Trp-Pro dipeptide fragment is proposed, based on the presented nmr data and the results of our molecular mechanics modeling of low-energy conformers of the peptides, reported elsewhere. © 1993 John Wiley & Sons, Inc.  相似文献   

13.
The redox potentials of the oriented films of the wild-type, the E194Q-, E204Q- and D96N-mutated bacteriorhodopsins (bR), prepared by adsorbing purple membrane (PM) sheets or its mutant on a Pt electrode, have been examined. The redox potentials (V) of the wild-type bR were −470 mV for the 13-cis configuration of the retinal Shiff base in bR and −757 mV for the all-trans configuration in H2O, and −433 mV for the 13-cis configuration and −742 mV for the all-trans configuration in D2O. The solvent isotope effect (ΔV=V(D2O)−V(H2O)), which shifts the redox potential to a higher value, originates from the cooperative rearrangements of the extensively hydrogen-bonded water molecules around the protonated CN part in the retinal Schiff base. The redox potential of bR was much higher for the 13-cis configuration than that for the all-trans configuration. The redox potentials for the E194Q mutant in the extracellular region were −507 mV for the 13-cis configuration and −788 mV for the all-trans configuration; and for the E204Q mutant they were −491 mV for the 13-cis configuration and −769 mV for the all-trans configuration. Replacement of the Glu194 or Glu204 residues by Gln weakened the electron withdrawing interaction to the protonated CN bond in the retinal Schiff base. The E204 residue is less linked with the hydrogen-bonded network of the proton release pathway compared with E194. The redox potentials of the D96N mutant in the cytoplasmic region were −471 mV for the 13-cis configuration and −760 mV for the all-trans configuration which were virtually the same as those of the wild-type bR, indicating that the D to N point mutation of the 96 residue had no influence on the interaction between the D96 residue and the CN part in the Schiff base under the light-adapted condition. The results suggest that the redox potential of bR is closely correlated to the hydrogen-bonded network spanning from the retinal Schiff base to the extracellular surface of bR in the proton transfer pathway.  相似文献   

14.
Five new phenylpropanoid amides, including N-trans-feruloyl-N′-cis-feruloyl-cadaverine (1), N,N′-trans-diferuloyl-3-oxo-cadaverine (2), N-trans-feruloyl-N′-cis-feruloyl-3-hydroxy-cadaverine (3), N,N′-cis-diferuloyl-3-hydroxy-cadaverine (4), N-trans-p-coumaroyl-N′-trans-feruloyl-3-hydroxy-cadaverine (5), were isolated from Alisma orientalis together with four known analogues. Their structural elucidations were conducted by using 1D and 2D NMR and HRESIMS spectroscopic analyses. The isolated compounds were assayed for their inhibitory activities against HCE-2, anti-oxidant effects, and their protective effects on H2O2-induced damage in human dopaminergic neuroblastoma cells (SH-SY5Y). Compounds 3, 6, and 7 displayed moderate anti-oxidant activities with IC50 values in the range of 36.940.7 μM. Compound 5 showed significant protective activity, while compounds 1, 2, 4, 7, and 8 showed moderate protective activities.  相似文献   

15.
Described herein are proton nmr experiments on chemically modified derivatives of ribonuclease A designed to elucidate the origin of an exchangeable resonance, assigned previously to a histidine ring N proton that titrates between 11 to 13 ppm with a pKa of 6.1 in H2O solution. Histidines 48 and 105, which are distant from the active site, are eliminated as candidates for this resonance from inhibitor binding studies on the enzyme in acetate–water solutions. This exchangeable resonance titrates with modified pKa's and constant area over the above pH range in His-119-N1-carboxymethylated-RNase A and des-(121–124)-RNase A, thus eliminating the imidazole N3 proton in the His 119-Asp 121 hydrogen bond. In His-12-N1-carboxymethylated-RNase A, this resonance is also observable, but broadens on raising the pH above 7 and at elevated temperatures above neutrality. It exhibits a pH-independent chemical shift characteristic of the protonated state of histidine. On the basis of these findings, this exchangeable resonance, designated a, is assigned to the imidazole N1 proton of His 12, which is hydrogen-bonded to the carbonyl oxygen of Thr 45 in the crystal.  相似文献   

16.
The interactions of cis- and trans-diammineplatinum compounds with 5′-GMP and 5′-dGMP in dilute aqueous solution at neutral pH were investigated by 1H nmr. In addition to the 1:2 Pt nucleotide complexes cis- and trans-Pt(NH3)2(GMP)2, it was possible to study the formation of the 1:1 Pt-nucleotide complexes with either one coordinated water or chloride ion. At 5°C GMP reacts with a stoichiometric amount of cis-diaquodiammine-platinum to yield cis-Pt(NH3)2(GMP) (H2O) as a sole reaction product. From the present results it is concluded that such a complex may play an important role as the initial reaction product between antitumor compounds like cis-Pt(NH3)2Cl2 and guanine in DNA in living organisms. The coupling constant 3J(H(1′)-H(2′)) of the H(1′) sugar proton in cis-Pt(NH3)2(GMP)2 is temperature dependent, indicating a conformational change in the sugar moiety.  相似文献   

17.
《Inorganica chimica acta》2004,357(10):2818-2826
[{Rh(cod)Cl}2] (cod=1,5-cyclooctadiene) reacts with o-(diphenylphosphino)benzaldehyde (PPh2(o-C6H4CHO)) (Rh:P=1:1) in the presence of aromatic diamines or 8-aminoquinoline (NN) to give acylhydride [Rh(Cl)(H){PPh2(o-C6H4CO)}(NN)] species. The oxidative addition of PPh2(o-C6H4CHO) in the presence of (NN) and PPh3 gives cationic species [Rh(H){PPh2(o-C6H4CO)} (PPh3)(NN)]+ containing mutually trans phosphorus atoms. When (NN)=8-aminoquinoline, a mixture of two isomers is obtained. These isomers differ in the nitrogen cis to the hydride, amino or quinolinic. By using Rh:PPh2(o-C6H4CHO)=1:2 stoichiometric ratios, oxidative addition of one PPh2(o-C6H4CHO) and P-coordination of another PPh2(o-C6H4CHO) occurs. The aldehyde group undergoes then a condensation reaction with the coordinated amine to afford new PNN terdentate ligands, phosphine-amino-imine when (NN)=diamine or phosphine-diimine when (NN)=8-aminoquinoline. These reactions give selectively the corresponding complexes [Rh(H){PPh2(o-C6H4CO)}(PNN)]+ containing trans phosphorus atoms and the hydride cis to the new imino group. X-ray diffraction studies of the PNN complexes are reported.  相似文献   

18.
cis,trans-Fe(CO)2(PMe3)2(p-Y-C6H4)X [X=Br, Y=H (4a), MeO (4b), Cl (4c), F (4d), Me (4e); X=I, Y=H (5); X=Cl, Y=H (6)] and cis,trans-Fe(CO)2(PMe3)2(σ-CHCH2)X [X=Br (7); X=I (8); X=Cl (9)] are prepared by reacting dihalide complexes cis,trans,cis- Fe(CO)2(PMe3)2X2 [X=Br (1), X=I (2), X=Cl (3)] with Grignard reagents p-Y-C6H4-MgBr (Y=H, OMe, Cl, F, Me) or CH2CH-MgBr and with lithium reagents PhLi, CH2CH-Li. With both reagents, the reaction proceeds following two parallel pathways: one is the metallation reaction which yields alkyl derivatives, the other affords 17 electron complexes [Fe(CO)2(PMe3)2X] via monoelectron reductive elimination. The influence of the halides and organometallic reagents on the yield of the metallation reaction is discussed. The solution structure of the complexes is assigned on the basis of IR and 1H, 13C, 19F, 31P NMR spectra. The solid state structure of complexes 4a, 5 and 6 is determined by single crystal X-ray diffractometric methods.  相似文献   

19.
Both trans- and cis-[PtCl2(NH3)(L)] compounds have been synthesized, L representing either the imino ether HN=C(OMe)Me having a Z or E configuration at the C=N double bond, or the cyclic ligands and (compounds 14 for trans geometry and 58 for cis geometry, respectively). The cyclic ligands mimic the imino ether ligands but, differently from imino ethers, cannot undergo change of configuration. In a panel of human tumor cells, trans compounds inhibit growth much more than transplatin. Moreover, compound 1 in most cases is less active than 2, and 1 and 2 are less active than 3 and 4, respectively. For cis compounds with imino ethers, the activity is reduced (5) or unaffected (6) with respect to cisplatin. Moreover, unlike trans compounds, substitution of cyclic ligands (7, 8) for imino ethers (5, 6) generally decreases the activity. This determines, for compounds with cyclic ligands, an unusual inversion of the cis geometry requirement for activity of platinum(II) species. Importantly, 14 and 58 partially circumvent the multifocal cisplatin resistance of A2780cisR cells, and 14 also overcome resistance from reduced uptake of 41McisR cells. DNA interaction regioselectivity of 14 and 58 is not substantially modified with respect to transplatin and cisplatin. However, both imino ethers and cyclic ligands slow down the DNA interstrand cross-link reaction, (E)-HN=C(OMe)Me and decreasing also its extent. Therefore, DNA interaction of 14 and 58 appears to be characterized by persistent monoadducts (14), and by monoadducts and/or intrastrand cross-links structurally different from those of cisplatin (58). This study demonstrates that ligand configuration modulates the activity of both trans and cis compounds, and supports the development of platinum drugs based on their coordination chemistry to combat cisplatin resistance.F.P. Intini and A. Boccarelli contributed equally to this work  相似文献   

20.
 The analogy between H-bonded nucleobase pairs and their metalated analogues is extended to the hemiprotonated pair of 7,9-dimethylguanine (7,9-DimeG) and the Watson-Crick and reversed Watson-Crick pair between 7,9-dimethylguaninium (7,9-DimeGH+) and 1-methylcytosine (1-MeC). The crystal structure analyses of two model compounds, trans–[Pt(CH3NH2)2(7,9-DimeG-N1)2](NO3)2 (1) and trans–[Pt(NH3)2(1-MeC-N3)(7, 9-DimeG-N1)](PF6)2· 2.5 H2O (3a) are reported. Pt binding is through N1 of 7,9-DimeG and N3 of 1-MeC. In solution, 3a exists in a mixture with Watson-Crick and reversed Watson-Crick arrangements of the two bases, depending on solvent, concentration and anions. Received: 16 October 1996 / Accepted: 27 January 1997  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号